Hostname: page-component-745bb68f8f-l4dxg Total loading time: 0 Render date: 2025-02-11T01:02:45.482Z Has data issue: false hasContentIssue false

Characteristics of the resonant charge transfer in strongly coupled plasmas including quantum and shielding effects

Published online by Cambridge University Press:  09 February 2016

Woo-Pyo Hong
Affiliation:
Department of Electronics Engineering, Catholic University of Daegu, Hayang 38430, South Korea
Young-Dae Jung*
Affiliation:
Department of Applied Physics and Department of Bionanotechnology, Hanyang University, Ansan, Kyunggi-Do 15588, South Korea Department of Physics, Applied Physics, and Astronomy, Rensselaer Polytechnic Institute, 110 8th Street, Troy, NY 12180-3590, USA
*
Email address for correspondence: ydjung@hanyang.ac.kr
Rights & Permissions [Opens in a new window]

Abstract

The influence of quantum tunnelling and shielding on the resonant electron transfer process in strongly coupled plasmas is investigated. The screened atomic states and energy eigenvalues are employed to obtain the resonant electron transfer cross-section in strongly coupled plasmas. It is found that the classical resonant electron transfer cross-section increases with an increase of the ion-sphere radius. However, the energy-dependent quantum tunnelling resonant electron transfer cross-section is shown to decrease with increasing ion-sphere radius. It is demonstrated that an increase of the nuclear charge decreases the screening effect on the electron transfer cross-section while the quantum tunnelling effect enhances the resonant electron transfer cross-section in strongly coupled plasmas. In addition, it is shown that the effect of quantum tunnelling on the resonant electron transfer process decreases when both the collision energy and ion-sphere radius increase. The variation of shielding effect on the resonant electron transfer process in strongly coupled plasmas is also discussed.

Type
Research Article
Copyright
© Cambridge University Press 2016 

1. Introduction

The charge exchange or transfer process between a positive ion and a neutral atom has been of a great interest since it is one of the most fundamental atomic collision processes in many areas of physics such as astrophysics, atomic physics, chemical physics and plasma physics, and also has many applications in plasma diagnostics (Rapp & Francis Reference Rapp and Francis1962; Janev, Presnyakov & Shevelko Reference Janev, Presnyakov and Shevelko1985; Shevelko & Vainshtein Reference Shevelko and Vainshtein1993; Beyer, Kluge & Shevelko Reference Beyer, Kluge and Shevelko1997; Song & Jung Reference Song and Jung2003; Jung Reference Jung2005a ; Fridman Reference Fridman2008). It is shown that the Debye–Hückel theory describes the physical properties of low density plasmas and also corresponds to a pair correlation approximation. Hence the electron capture processes in weakly coupled plasmas have been extensively investigated using the screened interaction of the collision system described by the Debye–Hückel potential since the average interaction energy between charged particles is usually smaller than the average kinetic energy of a plasma particle (Baimbetov, Nurekenov & Ramazanov Reference Baimbetov, Nurekenov and Ramazanov1995). However, it is known that the influence of multiparticle correlation caused by the collective interactions should be taken into account with increasing density in a plasma. Hence, the interaction potential in strongly coupled plasmas would not be properly described by the standard Debye–Hückel theory owing to the strong collective effects of multiparticle interactions. It is shown that the ion-sphere model (Salpeter Reference Salpeter1954; Jung Reference Jung1999; Fujimoto Reference Fujimoto2004) has also played a crucial role in elucidating the properties of the strongly coupled plasma and is also found to be more suitable than the cutoff potential model. Hence, it would be expected that the charge transfer and exchange processes in strongly coupled plasmas are quite different from those in weakly coupled plasmas. If the electron transfer in the ion-neutral collision has no energy exchange in the collision system; this process is known as the resonant electron transfer process; otherwise is called the non-resonant electron transfer case (Fridman Reference Fridman2008; Fridman & Kennedy Reference Fridman and Kennedy2011). The resonant and non-resonant electron transfer processes have been extensively explored using various classical and the quantum mechanical methods depending on the physical states of the ion-neutral collision system (Janev et al. Reference Janev, Presnyakov and Shevelko1985; Fridman Reference Fridman2008; Fridman & Kennedy Reference Fridman and Kennedy2011). In recent years, the quantum tunnelling phenomena have been comprehensively explored in wide areas of modern physics such as in semiconductors and superconductors, as well as in individual atoms on metal surfaces in various nanostructures including single-wall carbon nanotubes, molecular transistors and quantum wires formed in a two-dimensional electron gas (Komnik & Gogolin Reference Komnik and Gogolin2003; Razavy Reference Razavy2003). It is also shown that the quantum tunnelling effect on the resonant electron transfer process enhances the electron transfer cross-section (Fridman Reference Fridman2008; Fridman & Kennedy Reference Fridman and Kennedy2011). However, the resonant electron transfer process including the quantum tunnelling effect in strongly coupled plasmas has not been investigated as yet. Thus, in this paper, we investigate the plasma shielding and quantum tunnelling effects on the charge transfer process in collisions of a positive ion and a neutral atom in strongly coupled plasmas described by the ion-sphere model with the energy eigenvalue and ionization energy obtained by the Rayleigh–Ritz variational technique (Mathews & Walker Reference Mathews and Walker1970) since the theoretical atomic spectroscopy is essential to the study of plasma parameters as well as collision dynamics.

This paper is composed as follows: in § 2, we discuss the charge transfer process in strongly coupled plasmas. In § 3, we obtain the energy expectation value and ionization energy by using the Rayleigh–Ritz variational technique including the non-thermal and plasma shielding effects. In addition, we obtain the quantum tunnelling resonant electron transfer cross-section between the neutral atom and the ion in strongly coupled plasmas. In § 4, we discuss the shielding and quantum tunnelling effects on the resonant electron transfer process in strongly coupled plasmas. Finally, the conclusions are given in § 5.

2. Electron transfer process in strongly coupled plasmas

Since the resonant electron transfer has no effect on the electronic energy in the reaction system, the interaction between a positive ion $\text{A}^{+}$ with nuclear charge $Z_{A}e$ and a neutral atom B with nuclear charge $Z_{B}e$ is expressed by $\text{A}^{+}+\text{B}\rightarrow \text{A}+\text{B}^{+}$ . The potential energy of an electron in the field of free ions can be found in an excellent work of Fridman & Kennedy (Reference Fridman and Kennedy2011). The physical concept of Debye–Hückel theory (Kleinstreuer Reference Kleinstreuer2010) as a cooperative screening phenomenon is no longer applicable in very dense plasmas since the probability of finding other charged particles in a Debye sphere is quite small and almost vanishes in strongly coupled plasmas. It has been shown that the Debye–Hückel model would not be reliable to represent the interaction potential in strongly coupled plasmas. It has also been shown that the ion-sphere model (Salpeter Reference Salpeter1954; Salzmann Reference Salzmann1998) in strongly coupled plasmas is equivalent to the Wigner–Seitz sphere in condensed matter theory since the ion sphere consists of a single ion and its surrounding negatively charged sphere. The screened Coulomb interaction potential $V_{SC}(r)$ between the electron and the ion with charge $Ze$ in strongly coupled plasmas would be represented by the following ion-sphere model (Jung & Jeong Reference Jung and Jeong1996; Jung Reference Jung2000):

(2.1) $$\begin{eqnarray}V_{SC}(r)=-\frac{Ze^{2}}{r}\left[1-\frac{r}{2R_{I}}\left(3-\frac{r^{2}}{R_{I}^{2}}\right)\right]{\it\theta}(R_{I}-r),\end{eqnarray}$$

where $r$ is the interparticle distance, ${\it\theta}(R_{I}-r)~(=1\text{ for }R_{I}\geqslant r;=0\text{ for }R_{I}<r)$ is the step function and $R_{I}~(=[3(Z-1)/4{\rm\pi}n_{e}]^{1/3})$ is the ion-sphere radius given by the electron density $n_{e}$ and change number $Z$ since the total charge within the ion sphere would be neutral. As shown in (2.1), it is found that the ion-sphere potential $V_{SC}(R_{I})$ and its first derivative $\text{d}V_{SC}(r)/\text{d}r|_{r=R_{I}}$ vanish at the surface of the ion sphere. An improved version of the ion-sphere model has been proposed both in the bound state and in free electron terms taking into account relativistic and exchange effects (Salzmann Reference Salzmann1998). However, the original version of the ion-sphere model has been extensively used to explore various physical processes in high-density and low-temperature plasmas (Fujimoto Reference Fujimoto2004). Hence, the screened potential energy $V_{IS}(r,R_{I})$ of an electron in the field of ions $\text{A}^{+}$ and $\text{B}^{+}$ in strongly coupled plasmas based on the ion-sphere model is represented by

(2.2) $$\begin{eqnarray}\displaystyle V_{IS}(r,R_{I})=-\frac{e^{2}}{r}\left[1-\frac{r}{2R_{I}}\left(3-\frac{r^{2}}{R_{I}^{2}}\right)\right]-\frac{e^{2}}{|r_{AB}-r|}\!\left[1-\frac{|r_{AB}-r|}{2R_{I}}\left(3-\frac{|r_{AB}-r|^{2}}{R_{I}^{2}}\right)\!\right]\!, & & \displaystyle \nonumber\\ \displaystyle & & \displaystyle\end{eqnarray}$$

where $r$ is the distance between the electron and the ion $\text{A}^{+}$ and $r_{AB}$ is the interatomic distance. Since the ion-sphere radius $R_{I}$ is usually greater than the interatomic distance $r_{AB}$ in strongly coupled plasmas, the maximum potential energy $V_{IS}^{Max}$ would be obtained by the condition $\partial V_{IS}(r,R_{I})/\partial r=0$ with the constraint of $r<R_{I}$ as

(2.3) $$\begin{eqnarray}V_{IS}^{Max}(r=r_{AB}/2,R_{I})=-\frac{4e^{2}}{r_{AB}}\left[1-\frac{r_{AB}}{4R_{I}}\left(3-\frac{r_{AB}^{2}}{4R_{I}^{2}}\right)\right],\end{eqnarray}$$

where $V_{IS}^{Max}(r=r_{AB}/2,R_{I})$ is the maximum potential value at $r=r_{AB}/2$ . It should be understood that the classical analogy of the electron transfer process would be possible if the maximum potential energy $V_{IS}^{Max}$ were smaller than the initial energy $E_{B}$ as follows:

(2.4) $$\begin{eqnarray}\displaystyle E_{B}(Z_{B},r_{AB},R_{I}) & = & \displaystyle -\frac{I_{B}(Z_{B},R_{I})}{n^{2}}-\frac{e^{2}}{r_{AB}}\left[1-\frac{r_{AB}}{2R_{I}}\left(3-\frac{r_{AB}^{2}}{R_{I}^{2}}\right)\right]\nonumber\\ \displaystyle & {\geqslant} & \displaystyle -\frac{4e^{2}}{r_{AB}}\left[1-\frac{r_{AB}}{4R_{I}}\left(3-\frac{r_{AB}^{2}}{4R_{I}^{2}}\right)\right],\end{eqnarray}$$

where $I_{B}(Z_{B},R_{I})/n^{2}$ is the ionization energy of the electron in the neutral atom B with principal quantum number $n$ .

3. Atomic states and transfer cross-section

In strongly coupled plasmas, the radial Schrödinger equation (Jung Reference Jung2000) of the neutral atom B with the ion-sphere model and the effective screened charge $Z_{B}-{\it\delta}_{nl}(Z_{B},R_{I})$ is represented by

(3.1) $$\begin{eqnarray}-\frac{\hslash ^{2}}{2m}\left[\frac{1}{r^{2}}\frac{\text{d}}{\text{d}r}\left(r^{2}\frac{\text{d}R_{nl}(r)}{\text{d}r}\right)-\frac{l(l+1)}{r^{2}}R_{nl}(r)\right]-\frac{[Z_{B}-{\it\delta}_{nl}(Z_{B},R_{I})]e^{2}}{r}R_{nl}(r)=E_{nl}R_{nl}(r),\end{eqnarray}$$

where $\hslash$ is the rationalized Planck constant, $m$ is the mass of the electron and $R_{nl}(r)$ and $E_{nl}$ are the screened radial wave function and energy eigenvalue of the $nl$ th state, respectively. The inner screening constant ${\it\delta}_{nl}(Z_{B},R_{I})$ is determined by ${\it\delta}_{nl}(Z_{B},R_{I})=4{\rm\pi}\int _{0}^{n^{2}a_{0}/Z_{nl}}\text{d}r\,r^{2}n_{e}(R_{I})$ , $a_{0}~(=\hslash ^{2}/me^{2})$ is the Bohr radius of the hydrogen atom and $Z_{nl}$ is the effective $Z$ value in the $nl$ th state. For free atoms, the effective $Z$ values for atomic electrons have been introduced by the modified Slater rules (Jung & Gould Reference Jung and Gould1991). Using the $1s$ variation parameter ${\it\xi}_{1s}$ in the ion-sphere model, i.e. the effective Bohr radius in strongly coupled plasmas, the radial part of the screened normalized variational ground state wave function (Jung Reference Jung2000) $R_{1s}(r,{\it\xi}_{1s})$ is assumed to be

(3.2) $$\begin{eqnarray}R_{1s}(r,{\it\xi}_{1s})=\frac{2}{{\it\xi}_{1s}^{3/2}}\exp \left(-\frac{r}{{\it\xi}_{1s}}\right).\end{eqnarray}$$

Using the Rayleigh–Ritz variational technique (Mathews & Walker Reference Mathews and Walker1970) with the ansatz $R_{1s}(r,{\it\xi}_{1s})$ (3.2) and the energy expectation value $\langle E_{1s}(R_{I},{\it\xi}_{1s})\rangle$ for the ground state, the variation parameter (Jung Reference Jung2000) ${\it\xi}_{1s}(Z_{B},R_{I})$ for $a_{Z_{B}}<R_{I}$ would be obtained by the minimization condition $\partial \langle E_{1s}(R_{I},{\it\xi}_{1s})\rangle /\partial {\it\xi}_{1s}=0$ :

(3.3) $$\begin{eqnarray}{\it\xi}_{1s}(Z_{B},R_{I})/a_{Z_{B}}=\frac{1}{1-{\it\delta}_{1s}(Z_{B},R_{I})/Z_{B}},\end{eqnarray}$$

where $a_{Z_{B}}\equiv a_{0}/Z_{B}$ and the $1s$ screening constant ${\it\delta}_{1s}(Z_{B},R_{I})$ is found to be

(3.4) $$\begin{eqnarray}{\it\delta}_{1s}(Z_{B},R_{I})=\frac{(Z_{B}-1)(a_{Z_{B}}/R_{I})^{3}}{1-3(1-1/Z_{B})(a_{Z_{B}}/R_{I})^{3}}.\end{eqnarray}$$

The ionization energy (Jung Reference Jung2000) $I_{B}(Z_{B},R_{I})$ of the electron in the neutral atom B in strongly coupled plasmas is then represented by

(3.5) $$\begin{eqnarray}I_{B}(Z_{B},R_{I})=Z_{B}^{2}Ry[1-{\it\xi}_{1s}(Z_{B},R_{I})/a_{Z_{B}}]^{2},\end{eqnarray}$$

where $Ry~(=me^{4}/2\hslash ^{2}\approx 13.6~\text{eV})$ is the Rydberg constant. In (3.5), it is shown that the ionization limits would come down owing to the pressure ionization (Fujimoto Reference Fujimoto2004). After some mathematical manipulations using (2.3), (2.4), and (3.5), the upper limit of the screened interparticle distance $r_{AB}(Z_{B},R_{I})$ would be obtained by the following relation:

(3.6) $$\begin{eqnarray}r_{AB}(Z_{B},R_{I})\leqslant \frac{3e^{2}}{\displaystyle \frac{Z_{B}^{2}Ry[1-{\it\xi}_{1s}(Z_{B},R_{I})/a_{Z_{B}}]^{2}}{n^{2}}+\frac{3e^{2}}{2R_{I}}}.\end{eqnarray}$$

As seen in (3.6), the term $3e^{2}/2R_{I}$ in the denominator represents the shielding effect on the interparticle distance $r_{AB}(Z_{B},R_{I})$ in strongly coupled plasmas. Since the maximum screened interparticle distance $r_{AB}^{max}(Z_{B},R_{I})$ is represented by $r_{AB}^{max}(Z_{B},R_{I})=3e^{2}/[I_{B}(Z_{B},R_{I})/n^{2}+3e^{2}/2R_{I}]$ , the classical expression of the resonant electron transfer cross-section ${\it\sigma}_{Transf}^{Class}(Z_{B},R_{I})$ between the ion $\text{A}^{+}$ and the neutral atom B in strongly coupled plasmas is then found to be

(3.7) $$\begin{eqnarray}{\it\sigma}_{Transf}^{Class}(Z_{B},R_{I})=\frac{9{\rm\pi}e^{4}n^{4}/Z_{B}^{4}Ry^{2}}{\left[\displaystyle 1+3\frac{n^{2}}{Z_{B}}\left(\frac{a_{Z_{B}}}{R_{I}}\right)-2\left(1-\frac{1}{Z_{B}}\right)\left(\frac{a_{Z_{B}}}{R_{I}}\right)^{3}\right]^{2}},\end{eqnarray}$$

where the terms $3(n^{2}/Z_{B})(a_{Z_{B}}/R_{I})-2(1-1/Z_{B})(a_{Z_{B}}/R_{I})^{3}$ in the denominator represent the shielding effects on the classical resonant electron transfer cross-section. Hence, the classical characteristic function $C_{IS}(Z_{B},R_{I})$ for the influence of plasma shielding on the classical expression of the resonant electron transfer cross-section in strongly coupled plasmas based on the ion-sphere model is obtained by

(3.8) $$\begin{eqnarray}C_{IS}(Z_{B},R_{I})=\left[1+3\frac{n^{2}}{Z_{B}}\left(\frac{a_{Z_{B}}}{R_{I}}\right)-2\left(1-\frac{1}{Z_{B}}\right)\left(\frac{a_{Z_{B}}}{R_{I}}\right)^{3}\right]^{-2}.\end{eqnarray}$$

As shown in (3.8), it is found that the plasma shielding effect diminishes the classical expression of the resonant electron transfer cross-section in strongly coupled plasmas. When $Z_{B}\gg 1$ , from (3.7) and (3.8), the classical expression of the resonant electron transfer cross-section is found to be ${\it\sigma}_{Transf}^{Class}\propto Z_{B}^{4}$ . In the quantum mechanical charge transfer process, it has been shown that potential barrier penetration or the quantum tunnelling effect (Fridman Reference Fridman2008; Fridman & Kennedy Reference Fridman and Kennedy2011) enhance the resonant electron transfer cross-section for a positive ion and a neutral atom collision case. Using the Wentzel–Kramers–Brillouin (WKB) method (Razavy Reference Razavy2003), the electron tunnelling probability can be represented by the following exponential integral expression: $|T_{Tunnel}^{Q}|^{2}=\exp [-(2/\hslash )\int _{z_{1}}^{z_{2}}\,\text{d}z\sqrt{2m(V(z)-E)}]$ , where $z_{1}$ and $z_{2}$ are the turning points of the potential $V(z)$ . Using the screened potential energy $V_{IS}(r,R_{I})$ (2.2) with (3.3) and (3.5), the quantum tunnelling probability for the resonant electron transfer process between the ion $\text{A}^{+}$ and the neutral atom B in strongly coupled plasmas would be then obtained by

(3.9) $$\begin{eqnarray}|T_{Tunnel}^{Q}(Z_{B},R_{I},d)|^{2}\approx \exp \left[-\frac{2dZ_{B}}{\hslash }\sqrt{2mRy[1-{\it\xi}_{1s}(Z_{B},R_{I})/a_{Z_{B}}]}\right],\end{eqnarray}$$

where $d$ is the width of the potential barrier. Since the maximum width [7] $d_{Max}$ of the potential barrier would be determined when the electron tunnelling frequency exceeds the inverse ion-neutral atom collision time: $I_{B}(Z_{B},R_{I})|T_{Tunnel}^{Q}(Z_{B},R_{I},d)|^{2}/\hslash >v/d$ , where $v$ is the relative collision velocity, the quantum tunnelling resonant electron transfer cross-section can be represented by the geometric relation ${\it\sigma}_{transf}^{Q\text{-}Tunnel}\approx {\rm\pi}\,d_{Max}^{2}$ . Hence, the quantum tunnelling resonant electron transfer cross-section ${\it\sigma}_{transf}^{Q\text{-}Tunnel}(Z_{B},R_{I},\bar{E})$ between the neutral atom B and the ion $\text{A}^{+}$ in strongly coupled plasmas is found to be

(3.10) $$\begin{eqnarray}\displaystyle {\it\sigma}_{transf}^{Q\text{-}Tunnel}(Z_{B},R_{I},\bar{E}) & = & \displaystyle \frac{{\rm\pi}a_{Z_{B}}^{2}}{4\left[\displaystyle 1-2\left(1-\frac{1}{Z_{B}}\right)\left(\frac{a_{Z_{B}}}{R_{I}}\right)^{3}\right]}\nonumber\\ \displaystyle & & \displaystyle \times \,\left\{\ln \left[\frac{Z_{B}^{2}}{\sqrt{8\bar{E}}}\left(1-2\left(1-\frac{1}{Z_{B}}\right)\left(\frac{a_{Z_{B}}}{R_{I}}\right)^{3}\right)\right]\right\}^{2},\end{eqnarray}$$

where $\bar{E}\equiv (\hslash v/dRy)^{2}/8[=(\tilde{E}/2\bar{d}^{2})(m/{\it\mu})]$ , $\tilde{E}\equiv {\it\mu}v^{2}/2Ry$ , ${\it\mu}$ is the reduced mass of the colliding system $(\text{A}^{+},\text{B})$ and $\bar{d}\equiv d/a_{0}$ . When $Z_{B}\gg 1$ , from (3.10), the quantum tunnelling resonant electron transfer cross-section is found to be ${\it\sigma}_{transf}^{Q\text{-}Tunnel}\propto [\ln Z_{B}^{2}+f(\bar{E})]^{2}$ , where $f(\bar{E})$ is the energy-dependent function. Since the quantum tunnelling process is resolved by the relative collision velocity of the system, the quantum characteristic function $Q_{IS}(Z_{B},\bar{R}_{I},\bar{E})$ for the plasma shielding effect on the quantum tunnelling resonant electron transfer cross-section in strongly coupled plasmas is then determined by

(3.11) $$\begin{eqnarray}Q_{IS}(Z_{B},\bar{R}_{I},\bar{E})=\frac{\ln \left[\displaystyle \frac{Z_{B}^{2}}{\sqrt{8\bar{E}}}\left(1-2\left(1-\frac{1}{Z_{B}}\right)\left(\frac{1}{\bar{R}_{I}}\right)^{3}\right)\right]}{\left[\displaystyle 1-2\left(1-\frac{1}{Z_{B}}\right)\left(\frac{1}{\bar{R}_{I}}\right)^{3}\right]\ln \left(\displaystyle \frac{Z_{B}^{2}}{\sqrt{8\bar{E}}}\right)},\end{eqnarray}$$

where $\bar{R}_{I}(\equiv R_{I}/a_{Z_{B}})$ is the scaled ion-sphere radius. Using (3.7) and (3.10), the quantum tunnelling effect on the resonant electron transfer cross-section can also be obtained by the following expression of the tunnelling characteristic function $T_{IS}$ :

(3.12) $$\begin{eqnarray}\displaystyle T_{IS}(Z_{B},\bar{R}_{I},\bar{E}) & = & \displaystyle \frac{Z_{B}^{2}}{36n^{4}}\frac{\left[\displaystyle 1+3\frac{n^{2}}{Z_{B}}\left(\frac{a_{Z_{B}}}{R_{I}}\right)-2\left(1-\frac{1}{Z_{B}}\right)\left(\frac{a_{Z_{B}}}{R_{I}}\right)^{3}\right]^{2}}{1-2\left(\displaystyle 1-\frac{1}{Z_{B}}\right)\left(\displaystyle \frac{1}{\bar{R}_{I}}\right)^{3}}\nonumber\\ \displaystyle & & \displaystyle \times \,\ln \left[\frac{Z_{B}^{2}}{\sqrt{8\bar{E}}}\left(1-2\left(1-\frac{1}{Z_{B}}\right)\left(\frac{1}{\bar{R}_{I}}\right)^{3}\right)\right].\end{eqnarray}$$

As seen, the influence of quantum tunnelling on the resonant electron transfer process could be explicitly investigated through the logarithmic term in (3.12). Recently, the resonant charge transfer processes in collisions between positive ions have been investigated in strongly coupled plasmas (Jung Reference Jung2005b ) as well as in weakly coupled Lorentzian plasmas (Hong & Jung Reference Hong and Jung2014). However, the quantum tunnelling effect has not been included in the investigation of the resonant electron transfer process in strongly coupled plasmas. Hence, (3.10)–(3.12) are reliable to explore the influence of quantum tunnelling on the resonant electron transfer cross-section in strongly coupled plasmas based on the ion-sphere model. In addition, extensive and excellent investigations for the effective interaction potentials have been carried out in dense semiclassical and quantum plasmas (Deutsch Reference Deutsch1977; Deutsch, Gombert & Minoo Reference Deutsch, Gombert and Minoo1978; Ramazanov, Dzhumagulova & Gabdullin Reference Ramazanov, Dzhumagulova and Gabdullin2010; Ramazanov et al. Reference Ramazanov, Moldabekov, Dzhumagulova and Muratov2011; Shukla & Eliasson Reference Shukla and Eliasson2011; Akbari-Moghanjoughi & Shukla Reference Akbari-Moghanjoughi and Shukla2012; Shukla & Eliasson Reference Shukla and Eliasson2012; Akbari-Moghanjoughi Reference Akbari-Moghanjoughi2013; Dzhumagulova et al. Reference Dzhumagulova, Masheeva, Ramazanov and Donkó2014; Ramazanov et al. Reference Ramazanov, Moldabekov, Gabdullin and Ismagambetova2014). However, quantum tunnelling resonant electron transfer has not been explored as yet in semiclassical or quantum plasmas. Hence, the investigation on the quantum tunnelling and plasma shielding effects on the resonant electron transfer process in dense semiclassical and quantum plasmas will be treated elsewhere. It is important to note that the ion-sphere potential is designed so that potential and its first derivative with respect to $r$ vanish at the ion-sphere radius. In strongly coupled plasmas, plasma screening is better described by this ion-sphere picture, in which the stationary hydrogenic ion of total electric charge $Z-1$ is surrounded by $Z-1$ plasma electrons, uniformly distributed throughout the ion-sphere radius. Hence, the ion-sphere model would be reliable when the ion charge $Z$ is greater than unity since the ion-sphere radius $R_{I}\propto (Z-1)^{1/3}$ . In strongly coupled plasmas (Baimbetov, Nurekenov & Ramazanov Reference Baimbetov, Nurekenov and Ramazanov1996), the temperature $T$ and number density $n$ are shown to be approximately $(1{-}10)\times 10^{4}~\text{K}$ and $10^{19}{-}10^{20}\;\text{cm}^{-3}$ , respectively. Additionally, it has been shown that the physical properties of dense plasmas (Ramazanov et al. Reference Ramazanov, Dzhumagulova and Gabdullin2010) could be expressed by the plasma coupling parameter ${\it\Gamma}~[=(Ze)^{2}/ak_{B}T]$ , degeneracy parameter ${\it\theta}~(=k_{B}T/E_{F})$ and density parameter $r_{s}~(=a/a_{0})$ , where $E_{F}$ is the Fermi energy and $a$ is the average distance between particles in dense plasmas. Recently, the excellent works (Ramazanov et al. Reference Ramazanov, Dzhumagulova, Omarbakiyeva and Röpke2006; Omarbakiyeva, Ramazanov & Röpke Reference Omarbakiyeva, Ramazanov and Röpke2009) have provided the useful effective interaction potentials to describe electron–ion, electron–atom and ion–atom interactions in partially ionized dense plasmas taking into account quantum mechanical effects as well as plasma screening effects. In addition, an excellent investigation has provided the electron–atom interaction, including Pauli blocking and plasma screening in partially ionized hydrogen plasmas, using the Beth–Uhlenbeck approach (Omarbakiyeva et al. Reference Omarbakiyeva, Fortmann, Ramazanov and Röpke2010) to the second-virial coefficient.

Figure 1. The classical characteristic function $C_{IS}$ for the influence of plasma shielding on the classical expression of the resonant electron transfer cross-section in strongly coupled plasmas as a function of the scaled ion-sphere radius $\bar{R}_{I}$ for $n=1$ and $E_{F}/\hslash {\it\omega}_{p}=4$ , where ${\it\omega}_{p}$ is the electron plasma frequency. The solid line is the case of $Z=2$ and ${\it\Gamma}r_{s}{\it\theta}=7.32$ . The dashed line is the case of $Z=6$ and ${\it\Gamma}r_{s}{\it\theta}=65.08$ . The dotted line is the case of $Z=8$ and ${\it\Gamma}r_{s}{\it\theta}=111.12$ .

4. Discussion

In order to specifically investigate the plasma shielding and quantum tunnelling effects on the resonant electron transfer process, we consider the reaction: $\text{A}^{+}+\text{A}\rightarrow \text{A}+\text{A}^{+}$ , i.e. $\text{A}=\text{B}$ , then $a_{Z_{A}}=a_{Z_{B}}\equiv a_{Z}$ . Figure 1 represents the classical characteristic function $C_{IS}$ for the influence of plasma shielding on the classical expression of the resonant electron transfer cross-section as a function of the scaled ion-sphere radius $\bar{R}_{I}$ for various ion charge number $Z$ in strongly coupled plasmas. As shown in this figure, it is found that the classical characteristic function $C_{IS}$ increases with an increase of the ion-sphere radius $\bar{R}_{I}$ . Hence, we have found that the plasma shielding effect suppresses the classical resonant electron transfer cross-section in strongly coupled plasmas. It is also shown that the classical characteristic function $C_{IS}$ decreases with increasing ion charge number $Z$ . We then understand that the shielding effect on the classical resonant electron transfer cross-section is found to be more significant at larger ion charge numbers. Hence, we can expect that hydrogen or helium atom collisions would be less appropriate to investigate the plasma shielding effect in strongly coupled plasmas. Figure 2 shows the quantum characteristic function $Q_{IS}$ for the shielding effect on the quantum tunnelling resonant electron transfer cross-section as a function of the scaled ion-sphere radius $\bar{R}_{I}$ for various ion charge number $Z$ in strongly coupled plasmas. As seen, we find that the quantum characteristic function $Q_{IS}$ decreases with an increase of the ion-sphere radius $\bar{R}_{I}$ , which indicates that the shielding effect increases the quantum resonant electron transfer cross-section in strongly coupled plasmas. It is also shown that the quantum characteristic function $Q_{IS}$ increases with increasing ion charge number $Z$ . Hence, the shielding effect on the quantum resonant electron transfer cross-section is also more significant at larger ion charge numbers. Figure 3 represents the surface plot of the quantum characteristic function $Q_{IS}$ as a function of the scaled ion-sphere radius $\bar{R}_{I}$ and the scaled collision energy $\bar{E}$ , which demonstrates that the shielding effect on the quantum resonant electron transfer process decreases with increasing collision energy as well as ion-sphere radius. On the other hand, the energy dependence of the quantum characteristic function $Q_{IS}$ is found to be more significant at small ion-sphere radii. Figure 4 represents the tunnelling characteristic function $T_{IS}$ for the influence of quantum tunnelling on the resonant electron transfer cross-section as a function of the scaled ion-sphere radius $\bar{R}_{I}$ for various values of the collision energy $\bar{E}$ in strongly coupled plasmas. We note in figure 4 that the influence of quantum tunnelling strongly enhances the resonant electron transfer cross-section and decreases with an increase of the ion-sphere radius $\bar{R}_{I}$ . It is also found that the quantum tunnelling effect on the resonant electron transfer process decreases with increasing collision energy $\bar{E}$ in strongly coupled plasmas.

Figure 2. The quantum characteristic function $Q_{IS}$ for the shielding effect on the quantum tunnelling resonant electron transfer cross-section in strongly coupled plasmas as a function of the scaled ion-sphere radius $\bar{R}_{I}$ for $\bar{E}=0.001$ and $E_{F}/\hslash {\it\omega}_{p}=4$ . The solid line is the case of $Z=2$ and ${\it\Gamma}r_{s}{\it\theta}=7.32$ . The dashed line is the case of $Z=6$ and ${\it\Gamma}r_{s}{\it\theta}=65.08$ . The dotted line is the case of $Z=8$ and ${\it\Gamma}r_{s}{\it\theta}=111.12$ .

Figure 3. The surface plot of the quantum characteristic function $Q_{IS}$ as a function of the scaled ion-sphere radius $\bar{R}_{I}$ and the scaled collision energy $\bar{E}$ for $Z=8$ , $E_{F}/\hslash {\it\omega}_{p}=4$ and ${\it\Gamma}r_{s}{\it\theta}=111.12$ .

Figure 4. The tunnelling characteristic function $T_{IS}$ for the influence of quantum tunnelling on the resonant electron transfer cross-section in strongly coupled plasmas as a function of the scaled ion-sphere radius $\bar{R}_{I}$ for $n=1$ and $E_{F}/\hslash {\it\omega}_{p}=4$ , $Z=6$ , ${\it\Gamma}r_{s}{\it\theta}=65.08$ . The solid line is the case of $\bar{E}=0.001$ . The dashed line is the case of $\bar{E}=0.002$ . The dotted line is the case of $\bar{E}=0.003$ .

5. Conclusions

In this work we have investigated the shielding and quantum tunnelling effects on the resonant electron transfer process in strongly coupled plasmas based on the ion-sphere model. We also obtained the screened atomic states and energy eigenvalues by using the Rayleigh–Ritz variational method for the resonant electron transfer process in strongly coupled plasmas. In addition, we have derived an analytic expression for the resonant electron transfer cross-section in strongly coupled plasmas. In this work, we have found that the classical resonant electron transfer cross-section increases with an increase of the ion-sphere radius. However, we found that the energy dependent quantum tunnelling resonant electron transfer cross-section decreases with increasing ion-sphere radius. We have shown that the screening effect on the electron transfer cross-section decreases with an increase of the nuclear charge. We have also found that the quantum tunnelling effect enhances the resonant electron transfer cross-section in strongly coupled plasmas while the influence of quantum tunnelling on the resonant electron transfer process decreases with increasing collision energy and ion-sphere radius, as shown in figures 3 and 4. In conclusion, we have found that the shielding and quantum tunnelling effects play crucial roles on the resonant electron transfer process between a positive ion and a neutral atom in strongly coupled plasmas. These results would provide useful information on the plasma screening effect as well as the quantum mechanical tunnelling phenomena on the resonant charge exchange or transfer process in dense plasmas.

Acknowledgements

This paper is dedicated to the late Professor P. K. Shukla in memory of exciting and stimulating collaborations on physical processes in strongly coupled plasmas. One of the authors (Y.-D.J.) gratefully acknowledges Professor W. Roberge for useful discussions and warm hospitality while visiting the Department of Physics, Applied Physics, and Astronomy at Rensselaer Polytechnic Institute. This research was initiated while one of the authors (Y.-D.J.) was affiliated with Rensselaer as a visiting professor. One of the authors (Y.-D.J.) also acknowledges Professor T. Kato, Professor R. More, and Professor T. Fujimoto for useful discussions on the ion-sphere model while visiting the National Institute for Fusion Science, Japan. This research was supported by the Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education, Science and Technology (grant no. 2012-001493).

References

Akbari-Moghanjoughi, M. 2013 Shukla–Eliasson attractive force: revisited. J. Plasma Phys. 79, 189196.Google Scholar
Akbari-Moghanjoughi, M. & Shukla, P. K. 2012 Theory for large-amplitude electrostatic ion shocks in quantum plasmas. Phys. Rev. E 86, 066401.Google Scholar
Baimbetov, F. B., Nurekenov, Kh. T. & Ramazanov, T. S. 1995 Pseudopotential theory of classical non-ideal plasmas. Phys. Lett. A 202, 211.Google Scholar
Baimbetov, F. B., Nurekenov, K. H. T. & Ramazanov, T. S. 1996 Electrical conductivity and scattering sections of strongly coupled hydrogen plasmas. Physica A 226, 181.CrossRefGoogle Scholar
Beyer, H. F., Kluge, H.-J. & Shevelko, V. P. 1997 X-ray Radiation of Highly Charged Ions. Springer.Google Scholar
Deutsch, C. 1977 Nodal expansion in a real matter plasma. Phys. Lett. A 60, 317.Google Scholar
Deutsch, C., Gombert, M. M. & Minoo, H. 1978 Classical modelization of symmetry effects in the dense high-temperature electron gas. Phys. Lett. A 66, 381.CrossRefGoogle Scholar
Dzhumagulova, K. N., Masheeva, R. U., Ramazanov, T. S. & Donkó, Z. 2014 Effect of magnetic field on the velocity autocorrelation and the caging of particles in two-dimensional Yukawa liquids. Phys. Rev. E 89, 033104.CrossRefGoogle ScholarPubMed
Fridman, A. 2008 Plasma Chemistry. Cambridge University Press.Google Scholar
Fridman, A. & Kennedy, L. A. 2011 Plasma Physics and Engineering, 2nd edn. Taylor and Francis.Google Scholar
Fujimoto, T. 2004 Plasma Spectroscopy. Oxford University Press.Google Scholar
Hong, W.-P. & Jung, Y.-D. 2014 Quantum tunneling resonant electron transfer process in Lorentzian plasmas. Phys. Plasmas 21, 083301.Google Scholar
Janev, R. K., Presnyakov, L. P. & Shevelko, V. P. 1985 Physics of Highly Charged Ions. Springer.Google Scholar
Jung, Y.-D. 1999 Orientation phenomena for electron-ion collisional excitations in strongly coupled plasmas. Eur. Phys. J. D 7, 249.Google Scholar
Jung, Y.-D. 2000 Plasma screening effects on inelastic Compton scattering of photons by hydrogenic ions in strongly coupled classical plasmas. Phys. Plasmas 7, 819.Google Scholar
Jung, Y.-D. 2005a Collective effects on the symmetric resonant charge transfer in partially ionized hydrogen plasma. Appl. Phys. Lett. 86, 021502.Google Scholar
Jung, Y.-D. 2005b Plasma screening effects on resonant charge transfer in strongly coupled plasmas. Europhys. Lett. 69, 753.Google Scholar
Jung, Y.-D. & Gould, R. J. 1991 Energies and wave functions for many-electron atoms. Phys. Rev. A 44, 111.Google Scholar
Jung, Y.-D. & Jeong, H.-D. 1996 Bremsstrahlung in electron-ion Coulomb scattering in strongly coupled plasma using the hyperbolic-orbit trajectory method. Phys. Rev. E 54, 1912.Google Scholar
Kleinstreuer, C. 2010 Modern Fluid Dynamics. Springer.Google Scholar
Komnik, A. & Gogolin, A. O. 2003 Resonant tunneling between Luttinger liquids: a solvable case. Phys. Rev. Lett. 90, 246403.Google Scholar
Mathews, J. & Walker, R. L. 1970 Mathematical Methods of Physics, 2nd edn. Addison-Wesley.Google Scholar
Omarbakiyeva, Y. A., Fortmann, C., Ramazanov, T. S. & Röpke, G. 2010 Cluster virial expansion for the equation of state of partially ionized hydrogen plasma. Phys. Rev. E 82, 026407.Google Scholar
Omarbakiyeva, Y. A., Ramazanov, T. S. & Röpke, G. 2009 The electron–atom interaction in partially ionized dense plasmas. J. Phys. A 42, 214045.CrossRefGoogle Scholar
Ramazanov, T. S., Dzhumagulova, K. N. & Gabdullin, M. T. 2010 Effective potentials for ion-ion and charge-atom interactions of dense semiclassical plasma. Phys. Plasmas 17, 042703.CrossRefGoogle Scholar
Ramazanov, T. S., Dzhumagulova, K. N., Omarbakiyeva, Yu. A. & Röpke, G. 2006 Effective polarization interaction potentials of the partially ionized dense plasma. J. Phys. A 39, 4369.Google Scholar
Ramazanov, T. S., Moldabekov, Zh. A., Dzhumagulova, K. N. & Muratov, M. M. 2011 Pseudopotentials of the particles interactions in complex plasmas. Phys. Plasmas 18, 103705.Google Scholar
Ramazanov, T. S., Moldabekov, Zh. A., Gabdullin, M. T. & Ismagambetova, T. N. 2014 Interaction potentials and thermodynamic properties of two component semiclassical plasma. Phys. Plasmas 21, 012706.Google Scholar
Rapp, D. & Francis, W. E. 1962 Charge exchange between gaseous ions and atoms. J. Chem. Phys. 37, 2631.Google Scholar
Razavy, M. 2003 Quantum Theory of Tunneling. World Scientific.Google Scholar
Salpeter, E. E. 1954 Electrons screening and thermonuclear reactions. Austral. J. Phys. 7, 373.CrossRefGoogle Scholar
Salzmann, D. 1998 Atomic Physics in Hot Plasmas. Oxford University Press.Google Scholar
Shevelko, V. P. & Vainshtein, L. A. 1993 Atomic Physics for Hot Plasmas. Institute of Physics.Google Scholar
Shukla, P. K. & Eliasson, B. 2011 Colloquium: nonlinear collective interactions in quantum plasmas with degenerate electron fluids. Rev. Mod. Phys. 83, 885.Google Scholar
Shukla, P. K. & Eliasson, B. 2012 Novel attractive force between ions in quantum plasmas. Phys. Rev. Lett. 108, 165007.Google ScholarPubMed
Song, M.-Y. & Jung, Y.-D. 2003 Collective and plasma screening effects on electron transfer processes in nonideal plasmas. Phys. Scr. 68, 160.Google Scholar
Figure 0

Figure 1. The classical characteristic function $C_{IS}$ for the influence of plasma shielding on the classical expression of the resonant electron transfer cross-section in strongly coupled plasmas as a function of the scaled ion-sphere radius $\bar{R}_{I}$ for $n=1$ and $E_{F}/\hslash {\it\omega}_{p}=4$, where ${\it\omega}_{p}$ is the electron plasma frequency. The solid line is the case of $Z=2$ and ${\it\Gamma}r_{s}{\it\theta}=7.32$. The dashed line is the case of $Z=6$ and ${\it\Gamma}r_{s}{\it\theta}=65.08$. The dotted line is the case of $Z=8$ and ${\it\Gamma}r_{s}{\it\theta}=111.12$.

Figure 1

Figure 2. The quantum characteristic function $Q_{IS}$ for the shielding effect on the quantum tunnelling resonant electron transfer cross-section in strongly coupled plasmas as a function of the scaled ion-sphere radius $\bar{R}_{I}$ for $\bar{E}=0.001$ and $E_{F}/\hslash {\it\omega}_{p}=4$. The solid line is the case of $Z=2$ and ${\it\Gamma}r_{s}{\it\theta}=7.32$. The dashed line is the case of $Z=6$ and ${\it\Gamma}r_{s}{\it\theta}=65.08$. The dotted line is the case of $Z=8$ and ${\it\Gamma}r_{s}{\it\theta}=111.12$.

Figure 2

Figure 3. The surface plot of the quantum characteristic function $Q_{IS}$ as a function of the scaled ion-sphere radius $\bar{R}_{I}$ and the scaled collision energy $\bar{E}$ for $Z=8$, $E_{F}/\hslash {\it\omega}_{p}=4$ and ${\it\Gamma}r_{s}{\it\theta}=111.12$.

Figure 3

Figure 4. The tunnelling characteristic function $T_{IS}$ for the influence of quantum tunnelling on the resonant electron transfer cross-section in strongly coupled plasmas as a function of the scaled ion-sphere radius $\bar{R}_{I}$ for $n=1$ and $E_{F}/\hslash {\it\omega}_{p}=4$, $Z=6$, ${\it\Gamma}r_{s}{\it\theta}=65.08$. The solid line is the case of $\bar{E}=0.001$. The dashed line is the case of $\bar{E}=0.002$. The dotted line is the case of $\bar{E}=0.003$.